24
This is a repository copy of Modelling the feedbacks between mass balance, ice flow and debris transport to predict the response to climate change of debris-covered glaciers in the Himalaya. White Rose Research Online URL for this paper: http://eprints.whiterose.ac.uk/90105/ Version: Accepted Version Article: Rowan, A.V., Egholm, D.L., Quincey, D.J. et al. (1 more author) (2015) Modelling the feedbacks between mass balance, ice flow and debris transport to predict the response to climate change of debris-covered glaciers in the Himalaya. Earth and Planetary Science Letters, 430. 427 - 438. ISSN 0012-821X https://doi.org/10.1016/j.epsl.2015.09.004 [email protected] https://eprints.whiterose.ac.uk/ Reuse Unless indicated otherwise, fulltext items are protected by copyright with all rights reserved. The copyright exception in section 29 of the Copyright, Designs and Patents Act 1988 allows the making of a single copy solely for the purpose of non-commercial research or private study within the limits of fair dealing. The publisher or other rights-holder may allow further reproduction and re-use of this version - refer to the White Rose Research Online record for this item. Where records identify the publisher as the copyright holder, users can verify any specific terms of use on the publisher’s website. Takedown If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

Modelling the feedbacks between mass balance, ice flow and

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Modelling the feedbacks between mass balance, ice flow and

This is a repository copy of Modelling the feedbacks between mass balance, ice flow and debris transport to predict the response to climate change of debris-covered glaciers in theHimalaya.

White Rose Research Online URL for this paper:http://eprints.whiterose.ac.uk/90105/

Version: Accepted Version

Article:

Rowan, A.V., Egholm, D.L., Quincey, D.J. et al. (1 more author) (2015) Modelling the feedbacks between mass balance, ice flow and debris transport to predict the response to climate change of debris-covered glaciers in the Himalaya. Earth and Planetary Science Letters, 430. 427 - 438. ISSN 0012-821X

https://doi.org/10.1016/j.epsl.2015.09.004

[email protected]://eprints.whiterose.ac.uk/

Reuse

Unless indicated otherwise, fulltext items are protected by copyright with all rights reserved. The copyright exception in section 29 of the Copyright, Designs and Patents Act 1988 allows the making of a single copy solely for the purpose of non-commercial research or private study within the limits of fair dealing. The publisher or other rights-holder may allow further reproduction and re-use of this version - refer to the White Rose Research Online record for this item. Where records identify the publisher as the copyright holder, users can verify any specific terms of use on the publisher’s website.

Takedown

If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

Page 2: Modelling the feedbacks between mass balance, ice flow and

1

Modelling the feedbacks between mass balance, ice flow and debris 1

transport to predict the response to climate change of debris-covered 2

glaciers in the Himalaya 3

4

Ann V. Rowan1*, David L. Egholm2, Duncan J. Quincey3, Neil F. Glasser4 5

6

1Department of Geography, University of Sheffield, Sheffield, S10 2TN, UK 7

2Department of Geoscience, Aarhus University, Aarhus C, Denmark. 8

3School of Geography, University of Leeds, Leeds, LS2 9JT, UK 9

4Department of Geography and Earth Sciences, Aberystwyth University, Aberystwyth, SY23 10

3DB, UK 11

12

13

*Corresponding author: [email protected] 14

Page 3: Modelling the feedbacks between mass balance, ice flow and

2

Abstract 15

Many Himalayan glaciers are characterised in their lower reaches by a rock debris layer. This 16

debris insulates the glacier surface from atmospheric warming and complicates the response 17

to climate change compared to glaciers with clean-ice surfaces. Debris-covered glaciers can 18

persist well below the altitude that would be sustainable for clean-ice glaciers, resulting in 19

much longer timescales of mass loss and meltwater production. The properties and evolution 20

of supraglacial debris present a considerable challenge to understanding future glacier 21

change. Existing approaches to predicting variations in glacier volume and meltwater 22

production rely on numerical models that represent the processes governing glaciers with 23

clean-ice surfaces, and yield conflicting results. We developed a numerical model that 24

couples the flow of ice and debris and includes important feedbacks between debris 25

accumulation and glacier mass balance. To investigate the impact of debris transport on the 26

response of a glacier to recent and future climate change, we applied this model to a large 27

debris-covered Himalayan glacier—Khumbu Glacier in Nepal. Our results demonstrate that 28

supraglacial debris prolongs the response of the glacier to warming and causes lowering of 29

the glacier surface in situ, concealing the magnitude of mass loss when compared with 30

estimates based on glacierised area. Since the Little Ice Age, Khumbu Glacier has lost 34% 31

of its volume while its area has reduced by only 6%. We predict a decrease in glacier volume 32

of 8–10% by AD2100, accompanied by dynamic and physical detachment of the debris-33

covered tongue from the active glacier within the next 150 years. This detachment will 34

accelerate rates of glacier decay, and similar changes are likely for other debris-covered 35

glaciers in the Himalaya. 36

37

1. Introduction 38

Glaciers in the Himalaya are rapidly losing mass (Bolch et al., 2012). However, data 39

describing past and present glacier volumes are scarce, resulting in varying predictions of 40

future glacier volumes (Cogley, 2011; Kääb et al., 2012). To improve predictions of how 41

Himalayan glaciers will decline through the 21st Century and the impact on Asian water 42

resources, we need to quantify the processes that drive glacier change (e.g. Immerzeel et al., 43

2013; Pellicciotti et al., 2015; Ragettli et al., 2015; Shea et al., 2015). Changes in glacier 44

volume are driven by climate variations, particularly changes in atmospheric temperature and 45

precipitation amount, and modified by ice flow (Bolch et al., 2012; Kääb et al., 2012). The 46

lower portions of clean-ice glaciers lose mass rapidly during periods of warming. As glaciers 47

recede to higher elevations, a new equilibrium state between this smaller glacier and the 48

Page 4: Modelling the feedbacks between mass balance, ice flow and

3

warmer climate may be established. Numerical modelling is required to understand the 49

processes that cause glaciers to change because we cannot rely simply on the extrapolation of 50

present-day trends. Previous studies of Himalayan glaciers using models designed for clean-51

ice glaciers resulted in predictions of widespread rapid deglaciation (e.g. Shea et al., 2015). 52

However, debris-covered glaciers respond differently to warming because debris insulates the 53

ice surface (Jouvet et al., 2011; Kirkbride and Deline, 2013; Pellicciotti et al., 2015; Østrem, 54

1959). Debris-covered glaciers lose most mass by surface lowering rather than terminus 55

recession (Hambrey et al., 2008). Debris-covered glaciers can persist at lower elevations than 56

would be possible for an equivalent clean-ice glacier even when dramatically out of 57

equilibrium with climate (Anderson, 2000; Benn et al., 2012). As glaciers lose mass 58

preferentially from areas of clean ice and mass loss results in the melt-out of englacial debris, 59

debris coverage will increase as a glacier shrinks (Bolch et al., 2008; Kirkbride and Deline, 60

2013; Thakuri et al., 2014). Therefore, predicting the future of the Himalayan cryosphere and 61

water resources depends on understanding the impacts of climate change on debris-covered 62

glaciers. 63

64

Debris on glacier tongues is derived from surrounding hillslopes and is transported 65

englacially before resurfacing in the ablation zone (Fig. 1a). In times negative mass balance, 66

velocities decline and debris thickness at the ice surface increases (Kirkbride and Deline, 67

2013) (Fig. 1b). A thin layer of rock debris (0.01–0.1 m) enhances glacier surface ablation by 68

reducing albedo, whereas thicker rock debris reduces ablation by insulating the surface 69

(Mihalcea et al., 2008; Nicholson and Benn, 2006; Østrem, 1959). Thick supraglacial debris 70

causes a reversal of the mass balance gradient, with higher ablation rates upglacier than at the 71

terminus leading to reduced driving stresses and ice flow (Jouvet et al., 2011; Quincey et al., 72

2009). Spatial heterogeneity in debris thickness results in differential surface ablation and the 73

formation and decay of ice cliffs and supraglacial ponds that locally enhance ablation (Reid 74

and Brock, 2014; Sakai et al., 2000). An obstacle to understanding the behaviour of debris-75

covered glaciers lies in quantifying the highly variable distribution of debris across the 76

glacier surface and how this differs between glaciers. Supraglacial debris distribution and 77

thickness are difficult to determine remotely and laborious to measure directly (Mihalcea et 78

al., 2008; Nicholson and Benn, 2006; Reid et al., 2012), particularly over more than one 79

glacier (Pellicciotti et al., 2015). A further challenge to predicting the response of debris-80

covered glaciers to climate change is understanding not only the distribution of debris on a 81

glacier surfaces, but also how this varies over time. 82

Page 5: Modelling the feedbacks between mass balance, ice flow and

4

83

In the Himalaya, 14–18% of the total glacierised area is debris-covered (Kääb et al., 2012) 84

increasing to about 36% in the Everest region of Nepal which contains some of the longest 85

debris-covered glacier tongues in the world are found (Nuimura et al., 2012; Thakuri et al., 86

2014). Where debris cover on an individual glacier exceeds 40% of the total area mass loss is 87

mainly by terminus stagnation rather than recession (which requires a loss of mass whilst 88

maintaining flow towards the migrating terminus) (Immerzeel et al., 2013; Quincey et al., 89

2009). Some Himalayan glaciers are over 50% debris covered (Ragettli et al., 2015) and 90

debris is sufficiently thick to reduce rather than enhance ablation (Benn et al., 2012; Bolch et 91

al., 2008; Nicholson and Benn, 2006). In the Everest region of Nepal, 70% of the glacierised 92

area comprises just 40 of 278 glaciers, and these large glaciers are generally debris covered 93

(Thakuri et al., 2014) (Fig. 2a). Glaciers in the Everest region last advanced around 0.5 ka, a 94

period referred to as the Little Ice Age (LIA) but distinct from the European event of the 95

same name (Owen et al., 2009; Richards et al., 2000). Since the LIA, Everest-region glaciers 96

have consistently lost mass (Kääb et al., 2012; Nuimura et al., 2012). Between 1962 and 97

2011, the proportion of Everest region glaciers covered by rock debris has doubled due to 98

ongoing mass loss (Thakuri et al., 2014). 99

100

The future of debris-covered glaciers worldwide is uncertain due to the limitations of our 101

knowledge about the distribution of supraglacial debris and how this evolves over time. 102

Existing models designed for clean-ice glaciers or static assumptions that describe only the 103

present state of the glacier are difficult to extrapolate under a changing climate. Here, we use 104

a novel glacier model that includes the self-consistent development of englacial and 105

supraglacial debris and reproduces the feedbacks among mass-balance, ice-flow and debris 106

transport to investigate how debris modifies the behaviour of a Himalayan glacier in response 107

to climate change. As an example of how many debris-covered Himalayan glaciers respond 108

to climate change, we applied this model to the evolution of Khumbu Glacier in the Everest 109

region of Nepal from the Late Holocene advance (1 ka) to AD2200. 110

111

2. Khumbu Glacier, Nepal 112

Khumbu Glacier is a large debris-covered glacier in the Everest region (Fig. 2), with a length 113

of 15.7 km and area of 26.5 km2. The Changri Nup and Changri Shar Glaciers were 114

tributaries of Khumbu Glacier during the LIA but have since detached. The equilibrium line 115

altitude (ELA) estimated from mass balance measurements made in 1974 and 1976 is 5600 m 116

Page 6: Modelling the feedbacks between mass balance, ice flow and

5

(Benn and Lehmkuhl, 2000; Inoue, 1977; Inoue and Yoshida, 1980). More recent studies 117

have placed the ELA of Khumbu Glacier at 5700 m around AD2000 (Bolch et al., 2011) 118

within the icefall that links the accumulation area in the Western Cwm to the glacier tongue 119

(Fig. 3). The ELA may have increased due to atmospheric warming of about 0.9°C between 120

1994 and 2013 (Salerno et al., 2014). 121

122

The active part of Khumbu Glacier (the area exhibiting ice flow) receded towards the base of 123

the icefall since the end of the LIA while the total glacier length remained stable. Feature-124

tracking observations of velocities define the length of the active glacier as 10.3 km (62% of 125

the LIA glacier length) (Fig. 4). Decaying ice at the terminus beneath debris several metres 126

thick indicates terminus recession of less than 1 km since the LIA (Bajracharya et al., 2014). 127

We divide Khumbu Glacier into two parts based on observations of glacier dynamics; (1) the 128

active glacier where velocities range from 10 m to 70 m a-1 and mass is replenished from the 129

accumulation zone, and (2) the decaying tongue that no longer exhibits ice flow of more than 130

a few m a-1. Similar behaviour is reproduced by our glacier model and observed for many 131

glaciers in the Everest region (Quincey et al., 2009). 132

133

3. Methods 134

3.1 Bed topography 135

Ice thickness for Khumbu Glacier (Fig. 3) has been measured along seven transects down-136

glacier from the icefall using radio-echo sounding was 440 ± 20 m at 0.5 km below the icefall 137

close to Everest Base Camp, decreasing to less than 20 m at 4930 m at 2 km up-glacier of the 138

terminus (Gades et al., 2000). Gravity observations gave an ice thickness of 110 m adjacent 139

to Lobuche and 440 m adjacent to Gorak Shep (Moribayashi, 1978). No data exist above the 140

icefall. Ice thickness can be estimated by assuming that glacier ice behaves as a perfectly 141

plastic material such that thickness (h) is determined by surface slope (Į) and basal shear 142

stress (IJb) (Nye, 1952): 143

144

h = そ * (IJb / f * と * g * sin(g)) 145

146

where ȡ is the density of glacier ice, and g is acceleration due to gravity. A shape factor (f) 147

describes the aspect ratio of the cross-section of a valley glacier (Nye, 1952), and a down-148

glacier thinning factor (そ) describes the long profile: 149

150

Page 7: Modelling the feedbacks between mass balance, ice flow and

6

そ = 1 – a * xb 151

152

where a is a constant accounting for the length of the glacier, x is the flowline distance from 153

the headwall and b describes where thinning first occurs along the flowline. We estimated the 154

thickness of Khumbu Glacier at 35 regularly-spaced transects perpendicular to the central 155

flowline. Glacier topography was described using the ASTER GDEM 2011 Digital Elevation 156

Model (DEM) and the GLIMS outline (GLIMS et al., 2005). Values for IJb, f and そ were 157

determined by tuning against observations resulting in a mean IJb value of 150 kPa. Subglacial 158

bedrock topography was described by subtracting the interpolated ice thickness from the 159

DEM, smoothing and resampling to 100-m grid spacing. 160

161

3.2 Glacier topography 162

Topographic profiles were measured using a DEM with a 10-m grid spacing generated from 163

ALOS PRISM imagery acquired in 2006 (Fig. 2b). Glacier topography was calculated 164

perpendicular to the central flowline by taking the mean of a 200-m wide moving window. 165

The LIA glacier surface was reconstructed from the elevation of lateral and terminal moraine 166

crests which are preserved below the icefall (Fig. 2a). The LIA moraine crest was defined by 167

taking the maximum of a 300-m wide moving window centred on the moraine, and verified 168

in the field using a Garmin GPSmap 62s handheld unit (Fig 2c). There are no indicators of 169

past glacier topography above the icefall, so model simulations were fitted to the available 170

data from the ablation zone. 171

172

3.3 Glacier dynamics 173

Glacier velocities (i.e. surface displacements) were calculated using the panchromatic bands 174

of multi-temporal Landsat Operational Land Imager imagery and a Fourier-based cross-175

correlation feature tracking method (Luckman et al., 2007). The images were first co-176

registered with sub-pixel accuracy using large feature (128 x 128 pixels; 1920 m square) and 177

search (256 x 256 pixels; 3840 m square) windows focusing on non-glacierised areas. Glacier 178

displacements were then calculated using finer feature and search windows of 48 x 48 pixels 179

(720 m square) and 64 x 64 pixels (960 m square). Sufficiently robust correlations were 180

accepted on the strength of their signal-to-noise ratio and matches above an extreme 181

threshold of 100 m a-1 were removed as blunders. The remaining displacements were 182

converted to annual velocities assuming no seasonal variability in flow. Errors in the velocity 183

data comprise mismatches associated with changing surface features between images, and 184

Page 8: Modelling the feedbacks between mass balance, ice flow and

7

any inaccuracy in the image co-registration. Given that the glacier is slow-flowing (and thus 185

features do not change rapidly), and that the images were co-registered to a fraction of a 186

pixel, we estimate a maximum theoretical error of one pixel per year (i.e. 15 m). Empirically 187

measured displacements in stationary areas adjacent to the glacier suggest the real error is 188

around half this (i.e. 7–8 m a-1). 189

190

3.4 Numerical modelling 191

We used the ice model iSOSIA (Egholm et al., 2011) with a novel description of debris 192

transport that represents the self-consistent development of englacial and supraglacial debris 193

and reproduces the feedbacks amongst mass-balance, ice-flow and debris accumulation. 194

iSOSIA is a higher-order shallow-ice model, which in contrast to standard shallow-ice 195

approximation (SIA) models includes the effects of longitudinal and transverse stress 196

gradients. This makes iSOSIA more accurate than SIA models in settings where flow 197

velocities can vary over short distances, such as in steep and rugged terrains of alpine glaciers 198

(Egholm et al., 2011). Supraglacial debris across Himalayan glaciers is generally decimetres 199

to metres thick and acts to reduce rather than enhance ablation. Moreover, where debris cover 200

is thin in the upper part of the ablation zone of Khumbu Glacier, similar ablation rates are 201

observed for surfaces both with and without debris (Inoue and Yoshida, 1980). Therefore, 202

ablation beneath supraglacial debris was calculated using an exponential function that gave a 203

halving of ablation beneath 0.5 m of debris and assuming minimal ablation beneath a debris 204

layer with a thickness exceeding 1.0 m, in line with values calculated for neighbouring 205

Ngozumpa Glacier (Nicholson and Benn, 2006). 206

207

Transport of debris within and on top of the glacier was modelled as an advection problem 208

assuming that the ice passively transports the debris. Internal ice deformation and basal 209

sliding drive ice flow in iSOSIA and the depth-averaged flow velocity is therefore 210

211

四拍 噺 四拍鳥 髪 四長 212

213

The velocity due to ice deformation, 憲博鳥 ┸ is approximated as a tenth-order polynomial 214

function of ice thickness with coefficients that depend on ice surface slope and bed slope as 215

well as longitudinal stress and stress gradients (Egholm et al., 2011). 216

217

Page 9: Modelling the feedbacks between mass balance, ice flow and

8

Basal sliding is assumed to scale with the basal shear stress according to the following 218

empirical sliding model (Budd and Keage, 1979): 219

220 憲長 噺 稽鎚酵長陳軽勅

221

where 軽勅 is the effective pressure at the bed, and 稽鎚 噺 ね 抜 など貸替 m a-1 Pa-1 and 兼 噺 に are 222

constants. The basal shear stress is the bed-parallel stress vector at the base of the ice, which 223

is computed by projecting the full stress tensor at the base of the ice onto the glacier bed. The 224

shear stress is therefore sensitive to ice thickness, ice surface slope, local ice velocity 225

variation, as well as bed slope orientation (Egholm et al., 2011). The effective pressure was 226

assumed to be 20% of the ice overburden pressure. This standard approach (e.g. 227

Bindschadler, 1983; Braun et al., 1999; Egholm et al., 2012; Kessler et al., 2008) to 228

modelling basal sliding in alpine glaciers ignores the detailed distribution of water pressure as 229

well as ice-bed cavitation, which are both elements that we have no means of calibrating 230

empirically for Khumbu Glacier. We note that the distribution of sliding is thus considered 231

uncertain, also because the two sliding parameters 稽鎚 and 兼 are difficult to constrain 232

empirically; according to Budd et al. (1979), 兼 should vary between 1 and 3. On the other 233

hand, variations in sliding rate do not significantly influence our modelling results as long as 234

ice, and thus also englacial debris, is transported from the accumulation zone to the ablation 235

zone by either basal sliding or internal ice creep. 236

237

The debris concentration, c, at any point within the ice was updated through time, t, using the 238

following equation for debris advection: 239

240 項潔項建 噺 伐椛 ゲ 岶潔四岼 241

where u is the three-dimensional ice velocity vector. The equation is based on the assumption 242

that debris is transported passively with the ice, and hence that any change in debris 243

concentration in a point is controlled by the flux of debris and ice to and from that point. For 244

example, at the surface in the ablation zone, debris concentration generally increases over 245

time because melting of ice causes the total influx of ice by flow to be positive. Debris may 246

also accumulate along the base of the ice, because basal melting, controlled by the excess 247

Page 10: Modelling the feedbacks between mass balance, ice flow and

9

heat at the glacier bed (Egholm et al., 2012), drives ice towards the bed. However, most 248

debris follows a concave path from the ice surface in the accumulation zone, down to some 249

depth within the glacier, and then back to the glacier surface in the ablation zone. As a 250

boundary condition to the above equation, we assumed that debris is fed to the surface of the 251

glacier in the accumulation zone and that csa=0.001 (the concentration of debris at the ice 252

surface) is constant across the accumulation area. Debris in the high parts of Khumbu Glacier 253

is likely transported to the glacier by avalanches, and the high energy of the avalanches can 254

spread snow and debris across wide areas of the glacier surface. Without detailed knowledge 255

of the distribution and frequency of avalanches, we used a constant surface debris 256

concentration in the accumulation zone as the simplest possible boundary condition. We note, 257

however, that because localised quantities of debris in the accumulation zone have a tendency 258

of diffusing during transport in the glacier, the wide-spread distribution of debris near the 259

terminus of the glacier is largely insensitive to variations in the debris input distribution of 260

the accumulation zone. The order of csa was roughly estimated by considering the total area 261

of the surrounding ice-free hillslopes and assuming that the mean erosion rate is about 1 mm 262

a-1. The total hillslope sediment production was then uniformly distributed across the area of 263

the ice accumulation zone. We note that sediment production from these hillslopes varies 264

through time in response to variations in rock uplift and climate change (Scherler et al., 265

2011). However, our model experiments focus on the spatial patterns of debris distribution 266

and disregard any temporal evolution of debris production. The rate of debris input used here 267

should consequently only be regarded as a first-order estimate. 268

269

Debris transport was modelled using a three-dimensional grid. iSOSIA is a depth-integrated 270

2-D model, but for the purpose of tracking the three-dimensional debris transport, the 271

thickness of the ice was divided into 20 layers representing the vertical dimension of the 3-D 272

grid structure. iSOSIA only computes depth-averaged velocity components. However, to 273

capture velocity variations at depth within the ice we reconstructed in every time step the full 274

three-dimensional velocity field of the glacier. The vertical variation of velocity components 275

was derived from the assumption that the horizontal ice velocity caused by viscous ice 276

deformation decays as a fourth-order polynomial down through the ice, which is valid for 277

laminar flow of ice with a stress exponent of 3 (Van der Veen, 2013; p. 77). We calibrated 278

the fourth-order polynomial to yield the correct depth-averaged velocity: 279

280

Page 11: Modelling the feedbacks between mass balance, ice flow and

10

四岫権岻 噺 のね 釆な 伐 岾権月峇替挽 四拍 髪 四長

281

where 四拍 is the depth-averaged horizontal velocity and ub is basal sliding velocity. z is burial 282

depth below the ice surface and h is ice thickness. The internal vertical component of the ice 283

velocity, uv, was scaled linearly with accumulation/ablation at the surface (兼岌 鎚岻 and melting at 284

the glacier bed (兼岌 長岻: 285 憲塚岫権岻 噺 月 伐 権月 兼岌 鎚 髪 権月 兼岌 長

286

Melting at the bed is computed from the heat available at the bed: 287

288 兼岌 長 噺 圏長 髪 憲長 ゲ 酵長 伐 圏頂貢詣

289

where 圏長 噺 ど┻どねの W m-2 is the heat flux from the underlying crust; 憲長 ゲ 酵長 is the heat 290

produced at the bed by friction due to basal sliding; 貢 噺 ひぱど kg m-3 is the density of glacier 291

ice and 詣 噺 ぬぬね kJ kg-1 is the latent heat of ice. 圏頂 is the heat transported away from the 292

glacier bed by heat conduction in the overlying ice. It is estimated from the thermal gradient 293

at the glacier bed: 294 圏頂 噺 伐倦 項劇項権

295

and the thermal conductivity of ice, 倦 噺 に┻ね W m-1 K-1. The temperature field within the ice 296

was computed using the three-dimensional semi-implicit algorithm described by Egholm et 297

al. (2012). The rates of basal melting were typically limited to the order of 0.01 m a-1, which 298

is 1–2 orders of magnitude smaller than the rates of surface ablation. 299

300

The advection equation was integrated through time using explicit forward time stepping in 301

combination with a three-dimensional upwind finite-difference scheme. The size of the time 302

step was restricted by the Courant-Friedrichs-Lewy condition: 303

304 ッ建 判 なに ッmin憲max 305

Page 12: Modelling the feedbacks between mass balance, ice flow and

11

where ッmin is the smallest cell-dimension (along the x, y and z axes), and 憲max is the 306

maximum ice velocity component. Time steps were by this condition restricted to 1–5 model 307

days. The iSOSIA and debris transport algorithms were parallelised using OpenMP 308

(Chapman et al., 2007), and run on 12-core CPU servers. Each simulation typically lasted 8–309

12 hours. 310

311

3.5 Experimental design 312

Simulations were made for the catchment upstream from the base of the LIA terminal 313

moraine. The DEM was constructed from data collected between AD2001 and AD2010 so 314

we place the present day at the start of this window as AD2000. Mass balance was calculated 315

assuming linear temperature-dependent rates of accumulation and ablation following those 316

measured in 1974 and 1976 (Benn and Lehmkuhl, 2000; Inoue, 1977; Inoue and Yoshida, 317

1980). An atmospheric lapse rate of –0.004°C m-1 was calculated by linear regression of 318

MODIS Terra land surface temperatures (24/02/00–31/12/06) (NASA, 2001) for the Central 319

Himalayan region (Fig. 4). Glacier advance and recession were simulated by varying ELA 320

over time. Extreme topography results in the majority of glacier mass gain by avalanching 321

rather than direct snowfall, and the avalanche contribution to mass balance was estimated as 322

75% (Benn and Lehmkuhl, 2000). We removed snow and ice from slopes exceeding 28° and 323

redistributed the total volume uniformly on the accumulation area of the glacier surface. The 324

critical slope of 28° was selected because this threshold is low enough to prevent ice 325

accumulation on slopes that are clearly ice-free today, but high enough to not limit ice 326

accumulation on the glacier surface. 327

328

3.5.1 Initial Late Holocene simulation 329

Prior to the LIA (0.5 ka), Khumbu Glacier had a slightly greater extent during the Late 330

Holocene (~1 ka) and is likely to have reached the LIA extent by the formation of high 331

moraines that enclosed the glacier and drove the ice mass to thicken (Owen et al., 2009) (Fig. 332

2a). As a starting point for our transient simulations, we reconstructed the Late Holocene 333

glacier from an ice-free domain using an ELA of 5325 m over a 5000-year period. This 334

simulation was optimised to result in a steady-state glacier that provided a good fit to the Late 335

Holocene moraines (Fig. 5). A minor recession, inferred from the position of the LIA 336

moraines inside the Late Holocene moraines, was imposed after the Late Holocene advance 337

equivalent to an increase in ELA of 50 m to 5375 m over 500 years, and supraglacial debris 338

thickened due to the reduction in debris export as glacier velocities decreased. 339

Page 13: Modelling the feedbacks between mass balance, ice flow and

12

340

3.5.2 Simulation from the LIA to the present day 341

To simulate the LIA advance, maximum and recession, the ELA was increased from 5375 m 342

to 6000 m over 500 years. The distribution of englacial and supraglacial debris simulated for 343

the Late Holocene was used as a starting point for the LIA simulation. A range of present-day 344

ELA values (Fig. 3) was tested by comparing the simulated ice volume with observed glacier 345

topography; the best fit to the present-day ice thickness was an ELA of 6000 m. This places 346

the ELA of Khumbu Glacier at the top of the icefall rather than in the lower half as indicated 347

by recent measurements (Fig. 3). The simulated ice thicknesses were optimised to the LIA 348

moraines and the present-day glacier. This simulation ran to steady state to indicate how the 349

glacier would continue to evolve without any further change in climate. 350

351

3.5.3 Simulation from the present day to AD2200 352

Simulation of glacier change from the present day until AD2200 continued from the present-353

day simulation where the glacier was out of balance with climate. We imposed a linear rise in 354

ELA over 100 years from AD2000 to AD2100 equivalent to predicted minimum and 355

maximum warming relative to 1986–2005 by 2080–2099 of 0.9°C (increase in ELA of 225 m 356

assuming an atmospheric lapse rate of –0.004°C m-1) and 1.6°C (increase in ELA of 400 m), 357

in line with IPCC model ensemble predictions (CMIP5 RCP 4.5 scenario) (Collins et al., 358

2013). The simulation continued until AD2200 without any further change in climate. 359

360

3.6 Mass balance sensitivity 361

We tested the sensitivity of Khumbu Glacier to mass balance parameter values through the 362

LIA to the present day to assess the impact of these uncertainties on our projections for 363

AD2100. A range of present-day ELA values equivalent to a change in ELA of 150 m 364

(equivalent to ±0.3°C) produced a difference in glacier volume of 0.3 x 109 m3 (14% of 365

present-day volume) with no change in glacier length beyond the cell size of the model 366

domain (100 m). Lapse rates between –0.003°C m-1 and –0.006°C m-1 and no change in ELA 367

produced a difference in glacier volume of 0.4 x 109 m3 (19%) with no change in length. 368

Maintaining the relationship with temperature between rates of accumulation and ablation 369

whilst varying maximum values by ±10% produced a difference in glacier volume of 4.0 x 370

106 m3 (0.2%) with no change in length. 371

372

3.7 Comparison with simulations that do not transport debris 373

Page 14: Modelling the feedbacks between mass balance, ice flow and

13

To verify the effect of supraglacial debris on glacier change, the LIA to the present day was 374

simulated: (1) without the modification of ablation beneath the debris layer, that is, assuming 375

a clean rather than debris-covered surface, and (2) with maximum ablation reduced by 50% 376

(as in Section 3.6) to compare the impact of a uniform reduction in ablation, as sometimes 377

used when clean-ice glacier models are applied to debris-covered glaciers (Fig. 6). Mass loss 378

from the clean-ice glacier greatly exceeded that from the debris-covered glacier, resulting in a 379

glacier with 16% of the present-day volume and a 6.7 km reduction in length compared to the 380

dynamic debris-covered glacier simulated for the same period. A reduction in ablation of 381

50% resulted in dramatic mass loss to 27% of present-day volume and a 4.4 km reduction in 382

length compared to the dynamic debris-covered glacier simulated for the same period. Our 383

results highlight that the change in terminus position of debris-covered glaciers in response to 384

climate change is slower than for clean-ice glaciers. Similar behaviour is observed using 1-D 385

modelling (Banerjee and Shankar, 2013) and remote-sensing observations (Kääb et al., 2012). 386

Therefore, models developed for clean-ice glaciers using a uniform reduction in ablation do 387

not reliably simulate the evolution of debris-covered glaciers. 388

389

4. Results 390

4.1 Glacier morphology 391

Reconstruction of Khumbu Glacier using moraine crests showed that, since the LIA, glacier 392

area has decreased from 28.1 km2 to 26.5 km2 (a reduction of 6%). If the glacier is considered 393

only in terms of active ice, then area has declined to 20.3 km2 (a reduction of 28%) (Fig. 4). 394

These values exclude the change in area attributed to the dislocation of the Changri Nup and 395

Changri Shar tributaries (Fig. 4). The volume of the active glacier is 1.7 x 109 m3 (50% of the 396

LIA volume). The lack of dynamic behaviour in the tongue can be observed from the relict 397

landslide material on the true left of the glacier that has not moved between 2003 and 2014 398

(Fig. 2a). Comparison of swath topographic profiles of the glacier surface and the LIA lateral 399

moraine crests (Fig. 2c) indicate mean surface lowering across the debris-covered tongue of 400

25.5 ± 10.6 m, or 0.05 ± 0.02 m a-1 since the LIA. Glacier volume decreased from 3.4 x 109 401

m3 to 2.3 x 109 m3 (66% of the LIA volume), a loss of 1.2 x 109 m3 and equivalent to 2.3 x 106 402

m3 a-1. Mean surface lowering observed between 1970 and 2007 across the ablation area was 403

13.9 ± 2.5 m (Bolch et al., 2011) suggesting that rates of mass loss have accelerated over the 404

last 50 years compared to the last 500 years, and consistent with the observed decrease in the 405

active glacier area (Quincey et al., 2009). 406

407

Page 15: Modelling the feedbacks between mass balance, ice flow and

14

4.2 Glacier modelling 408

The initial simulation representing the Late Holocene maximum was computed from an ice-409

free domain using an ELA of 5325 m (–2.7°C relative to the present day). Debris 410

accumulated at the ice margins rather than on the glacier surface to form lateral moraines 411

(Fig. 5). 412

413

4.2.1 The Little Ice Age to the present day 414

Khumbu Glacier initially advanced during the LIA for 150 years despite the rise in ELA as 415

decreasing velocity in the tongue (Table 1) resulted in thickening supraglacial debris (Fig. 7a 416

and 7c). The large LIA moraines suggest that debris export from the glacier to the ice 417

margins declined because the glacier was impounded following the construction of these 418

moraines. This simulation reproduced this observation, and resulted in the formation of a 419

thick debris layer (Fig. 7d). The simulated LIA glacier surface provided a good fit to the LIA 420

moraine crests (Fig. 7e). The simulated glacier then lost mass by surface lowering 421

accompanied by minor terminus recession, despite the reduction in ablation beneath 422

supraglacial debris (Fig 7b and Table 1). Simulated present-day ice thicknesses were in good 423

agreement with the observed glacier surface (Fig. 7f). The maximum simulated present-day 424

ice thickness was 345 m. The mean flowline ice thickness was 168 m for the whole glacier, 425

88 m in the accumulation area and 210 m for the debris-covered tongue. Simulated velocities 426

(Table 1 and Fig. 8) reproduced the pattern and absolute values measured from remote-427

sensing observations (Fig. 4). 428

429

After the LIA maximum, simulated ice thickness declined most rapidly for the first 200 years 430

of warming followed by slightly less rapid mass loss for the following 300 years. Mean ice 431

thickness across the entire glacier decreased by 0.01 m a-1, and surface lowering was greatest 432

between 1.8 km and 3.2 km upglacier from the terminal moraine. The active glacier shrunk to 433

the observed active ice extent but did not reach steady state. The response time to reach 434

equilibrium with the present-day ELA was 1150 years, 500 years longer than the time elapsed 435

between the LIA maximum and the present day, indicating that Khumbu Glacier is out of 436

balance with climate. According to our model, Khumbu Glacier will continue to respond to 437

post-LIA warming until about AD2500 and will lose a further 0.4 x 109 km3 (18%) of ice 438

without any further change in climate. 439

440

4.2.2 Present day to AD2200 441

Page 16: Modelling the feedbacks between mass balance, ice flow and

15

To predict glacier volume at AD2100 and AD2200, we imposed a linear rise in ELA from the 442

present day following IPCC minimum and maximum warming scenarios for AD2100 443

(Collins et al., 2013). These simulations were driven by an increase in ELA of 225 m to 6225 444

m (equivalent to warming of 0.9°C) and 400 m to 6400 m (equivalent to warming of 1.6°C) 445

over 100 years, and without a further change in climate until AD2200. Warming of 0.9°C by 446

AD2100 will result in mass loss of 0.17 x 109 km3 and warming of 1.6°C will result in mass 447

loss of 0.21 x 109 km3 (Fig. 9a and 9c), a decrease in glacier volume of between 8% and 10% 448

(Table 1). Simulated mass loss will be greatest close to the base of the icefall, where ablation 449

exceeds that occurring down-glacier beneath thicker supraglacial debris and up-glacier in the 450

Western Cwm. Supraglacial debris will expand and thicken across the glacier tongue, 451

particularly between the confluence with Changri Nup Glacier and the icefall (Fig. 9e 452

compared to Fig. 7d), reaching 1.5 m thickness at the base of the icefall. The debris-covered 453

tongue could physically detach from the base of the icefall within 150 years and persist in 454

situ while the active glacier recedes (Fig. 9b and 9d). After the physical detachment of the 455

debris-covered tongue, supraglacial debris will develop on the tongue of the active glacier 456

near the upper part of the icefall (Fig. 9f). 457

458

5. Discussion 459

5.1 Validation of model simulations 460

The present-day simulation was validated by comparison with observations of velocities, 461

mean surface elevation change and geodetic mass balance derived from satellite imagery. The 462

simulated present-day maximum flowline velocity was 59 m a-1 and the mean was 9 m a-1 463

(Fig. 8a and 8b). The mean simulated velocity above the base of the icefall was 24 m a-1, and 464

the mean velocity of the debris-covered tongue below the icefall was 2 m a-1. These 465

simulated velocities are in good agreement with those measured using feature tracking (Fig. 466

4), which give a maximum flowline velocity of 67 m a-1 and a mean of 16 m a-1. The mean 467

measured velocity above the base of the icefall was 25 m a-1, and the mean velocity of the 468

debris-covered tongue was 9 m a-1. The measured velocity of the tongue is within the 469

uncertainty of the feature tracking method due to the 15-m grid spacing of the imagery used, 470

and the actual displacement could be less than 9 m a-1. 471

472

The decrease in the elevation of the simulated glacier surface over the 40 years prior to the 473

present day was close to zero at the terminus and increased to 8–10 m in the upper part of the 474

ablation area. The simulated surface lowering shows good agreement both in terms of the 475

Page 17: Modelling the feedbacks between mass balance, ice flow and

16

absolute values and the distribution of surface lowering to that observed for a similar period 476

(1970–2007) which gave an elevation difference of –13.9 ± 2.5 m across the ablation area 477

(Bolch et al., 2011). Integrated mass balance for the simulated present-day glacier was –0.22 478

m w.e. a-1, slightly less negative than but not dissimilar to geodetic mass balance values 479

estimated between 1970 and 2007 as of –0.27 ± 0.08 m w.e. a-1 (Bolch et al., 2011) and 480

between 1992 and 2008 as –0.45 ± 0.52 m w.e. a-1 (Nuimura et al., 2012). 481

482

5.2 Equilibrium Line Altitude 483

The ELA of Khumbu Glacier could be placed in a range from 5200 m to 5600 m assuming 484

that the integrated mass balance is zero (Benn and Lehmkuhl, 2000). However, methods for 485

calculating ELA such as the accumulation-area ratio are difficult to apply to avalanche-fed, 486

debris-covered glaciers for which values appear to be lower (around 0.1–0.4) than those for 487

clean-ice glaciers (0.5–0.6) (Anderson, 2000). Snowline altitude is not a reliable indicator of 488

ELA in high mountain environments, because avalanching, debris cover and high relief affect 489

mass balance such that ELA may differ by several hundred metres from the mean snowline 490

(Benn and Lehmkuhl, 2000). Simulations using the lower estimated ELAs and assuming a net 491

mass balance of zero produced a glacier equivalent to the Late Holocene extent. Simulations 492

optimised to the present-day glacier indicate that ELA is probably about 5800–6000 m (Fig. 493

3b). 494

495

5.3 Sources of uncertainty associated with modelling debris-covered glaciers 496

We used a simple approach to represent the relationship between climate and glacier mass 497

balance to avoid introducing additional uncertainties by making assumptions about the 498

response of meteorological parameters such as monsoon intensity to climate change. 499

Therefore, our results indicate the sensitivity of a debris-covered Himalayan glacier to 500

climate change over the Late Holocene period (1 ka to present). Although iSOSIA captures 501

the dynamics of mountain glaciers, the interaction of high topography with atmospheric 502

circulation systems will affect mass balance (Salerno et al., 2014). Future studies could use 503

downscaled climate model outputs or energy balance modelling to better capture these 504

variables. However, mass balance and meteorological data to support these approaches are 505

scarce for the majority of Himalayan glaciers. 506

507

Differences in the estimated and simulated volume of the present-day glaciers were due to 508

differences in simulated glacier extents. Simulations were designed to give a best fit to 509

Page 18: Modelling the feedbacks between mass balance, ice flow and

17

Khumbu Glacier and produced less extensive ice than observed for Changri Nup and Changri 510

Shar Glaciers (Fig. 7a and b). Sensitivity experiments showed that a range of mass balance 511

values and lapse rates had little impact on these tributaries, suggesting that the mass balance 512

of Khumbu Glacier does not precisely represent that of the tributaries. This mismatch could 513

be due to the differences in hypsometry between glaciers and model calibration to the 514

extreme altitudes in the Western Cwm. 515

516

There are no measurements with which to constrain ice thickness in the Western Cwm, so our 517

estimate of ice thickness is based solely on the slope of the glacier surface derived from the 518

DEM and tuning of values for basal shear stress (IJb) and glacier shape to match geophysical 519

observations (Fig. 3). The IJb values initially used to determine bed topography are within the 520

range simulated using iSOSIA (Fig. 8a and 8b) suggesting that our estimate of bed 521

topography is appropriate. However, calculation of bed topography beneath glaciers and ice 522

sheets remains an outstanding challenge in glaciology, and one that is difficult to resolve in 523

the absence of data describing the basal properties of the glacier. 524

525

The addition of debris to the glacier surface by rock avalanching from the surrounding 526

hillslopes is not represented in our glacier model, but previous studies have demonstrated that 527

large rock avalanches can perturb the terminus position of mountain glaciers (e.g. Menounos 528

et al., 2013; Vacco et al., 2010). Sub-debris ablation is modified by the physical properties of 529

the debris layer, particularly variations in water content and grain size (Collier et al., 2014). 530

Exposed ice cliffs can enhance ablation locally on debris-covered glaciers; at Miage Glacier 531

in the European Alps ice cliffs occupy 1% of the debris-covered area and account for 7% of 532

total ablation (Reid and Brock, 2014). Previous work has hypothesised that ice cliff ablation 533

may be responsible for the comparable rates of mass loss observed for debris-covered and 534

clean-ice glaciers in the Himalaya and Karakoram (Gardelle et al., 2012; Kääb et al., 2012). 535

Mapping the area of debris-covered surfaces occupied by ice cliffs and increasing ablation 536

accordingly could refine our predictions of future glacier change. This would require more 537

detailed topographic data than the 30-m DEM and parameterisation of the processes by which 538

ice cliffs form and decay. As we do not incorporate ice cliffs or supraglacial ponds into our 539

modelling, and as these features are likely to become more widespread as surface lowering 540

continues, our estimates of mass loss from the present day to AD2200 are likely to be 541

cautious. 542

543

Page 19: Modelling the feedbacks between mass balance, ice flow and

18

6. Conclusions 544

Predictions of debris-covered glacier change based either on assumptions about clean-ice 545

glaciers or including static adjustments of ablation rates do not capture the feedbacks 546

amongst mass balance, ice dynamics and debris transport that govern the behaviour of these 547

glaciers, and are unlikely to give reliable results. We present the first dynamic model of the 548

evolution of a debris-covered glacier and demonstrate that including these important 549

feedbacks simulated glacier mass loss by surface lowering rather than terminus recession, and 550

represents the observed response to climate change of debris-covered glaciers. Models such 551

as this that represent the transient processes governing the behaviour of debris-covered 552

glaciers, supported by detailed direct and remotely-sensed observations, are needed to 553

accurately predict glacier change in mountain ranges such as the Himalaya. 554

555

The development of supraglacial debris on Khumbu Glacier in Nepal promoted a reversed 556

mass balance profile across the ablation area, resulting in greatest mass loss after the Little 557

Ice Age (0.5 ka) where debris was absent close to the icefall and least mass loss on the 558

debris-covered tongue. The reduction in ablation across the debris-covered section of the 559

glacier resulted in reduced ice flow and debris export. Khumbu Glacier extends to a lower 560

altitude (4870 m a.s.l. compared to 5160 m a.s.l.) and greater length (15.7 km compared to 561

10.3 km) than would be possible without supraglacial debris. We predict a loss of ice 562

equivalent to 8–10% of the present-day glacier volume by AD2100 with only minor change 563

in glacier area and length, and physical detachment of the debris-covered tongue from the 564

upper active part of the glacier before AD2200. Regional atmospheric warming is likely to 565

result in a similar response from other debris-covered glaciers in the Everest region over the 566

same period. 567

568

569

570

Acknowledgements We thank S. Brocklehurst for critical discussion and reading of the 571

manuscript. Some of this research was undertaken while A.V.R. was supported by a Climate 572

Change Consortium of Wales (C3W) postdoctoral research fellowship at Aberystwyth 573

University. The glacier model simulations were performed on the Iceberg High-Performance 574

Computer at the University of Sheffield. ASTER GDEM is a product of METI and NASA. 575

Two anonymous reviewers are thanked for helpful comments that improved the clarity of this 576

manuscript. 577

Page 20: Modelling the feedbacks between mass balance, ice flow and

19

578

579

References 580

Anderson, R.S., 2000. A model of ablation-dominated medial moraines and the generation of 581

debris-mantled glacier snouts. J Glacio 46, 459–469. 582

Bajracharya, S.R., Maharjan, S.B., Shrestha, F., Bajacharya, O.M., Baidya, S., 2014. Glacier 583

Status in Nepal and Decadal Change from 1980 to 2010 Based on Landsat Data. 584

ICIMOD research report. 585

Banerjee, A., Shankar, R., 2013. On the response of Himalayan glaciers to climate change. J 586

Glacio 59, 480–490. doi:10.3189/2013JoG12J130 587

Benn, D., Benn, T., Hands, K., Gulley, J., Luckman, A., Nicholson, L.I., Quincey, D., 588

Thompson, S., Toumi, R., Wiseman, S., 2012. Response of debris-covered glaciers in the 589

Mount Everest region to recent warming, and implications for outburst flood hazards. 590

Earth Science Reviews 114, 156–174. doi:10.1016/j.earscirev.2012.03.008 591

Benn, D.I., Lehmkuhl, F., 2000. Mass balance and equilibrium-line altitudes of glaciers in 592

high-mountain environments. Quaternary International 65, 15–29. 593

Bindschadler, R.A., 1983. The importance of pressurized subglacial water in separation and 594

sliding at the glacier bed. J Glacio 29, 3–19. 595

Bolch, T., Buchroithner, M., Pieczonka, T., Kunert, A., 2008. Planimetric and volumetric 596

glacier changes in the Khumbu Himal, Nepal, since 1962 using Corona, Landsat TM and 597

ASTER data. J Glacio 54, 592–600. 598

Bolch, T., Kulkarni, A., Kääb, A., Huggel, C., PAUL, F., Cogley, J.G., Frey, H., Kargel, J.S., 599

Fujita, K., Scheel, M., Bajracharya, S., Stoffel, M., 2012. The State and Fate of 600

Himalayan Glaciers. Science 336, 310–314. doi:10.1126/science.1215828 601

Bolch, T., Pieczonka, T., Benn, D.I., 2011. Multi-decadal mass loss of glaciers in the Everest 602

area (Nepal Himalaya) derived from stereo imagery. The Cryosphere 5, 349–358. 603

doi:10.5194/tc-5-349-2011 604

Braun, J., Zwartz, D., Tomkin, J., 1999. A new surface-processes model combining glacial 605

and fluvial erosion. Annals of Glaciology 28, 282–290. 606

Budd, W., Keage, P., 1979. Empirical studies of ice sliding. J Glaciol 23, 157–170. 607

Chapman, B., Jost, G., van der Pas, R., 2007. Using OpenMP. The MIT Press. 608

Cogley, J.G., 2011. Present and future states of Himalaya and Karakoram glaciers. Ann. 609

Glaciol 52, 69–73. 610

Collier, E., Nicholson, L.I., Brock, B.W., Maussion, F., Essery, R., Bush, A.B.G., 2014. 611

Representing moisture fluxes and phase changes in glacier debris cover using a reservoir 612

approach. The Cryosphere 8, 1429–1444. doi:10.5194/tc-8-1429-2014 613

Collins, M., Knutti, R., Arblaster, J., 2013. Long-term Climate Change: Projections, Com- 614

mitments and Irreversibility. In: Climate Change 2013: The Physical Science Basis. 615

Contribution of Working Group I to the Fifth Assessment Report of the 616

Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. 617

Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. 618

Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA 1–619

108. 620

Egholm, D., Knudsen, M., Clark, C., Lesemann, J.E., 2011. Modeling the flow of glaciers in 621

steep terrains: The integrated second-order shallow ice approximation (iSOSIA). J. 622

Geophys. Res 116, F02012. 623

Egholm, D.L., Pedersen, V.K., Knudsen, M.F., Larsen, N.K., 2012. Coupling the flow of ice, 624

water, and sediment in a glacial landscape evolution model. Geomorphology 141-142, 625

47–66. doi:10.1016/j.geomorph.2011.12.019 626

Page 21: Modelling the feedbacks between mass balance, ice flow and

20

Gades, A., Conway, H., Nereson, N., Naito, N., Kadota, T., 2000. Radio echo-sounding 627

through supraglacial debris on Lirung and Khumbu Glaciers, Nepal Himalayas. IAHS 628

Publications 264, 13–24. 629

Gardelle, J., Berthier, E., Arnaud, Y., 2012. Slight mass gain of Karakoram glaciers in the 630

early twenty-first century. Nature Geosci 5, 1–4. doi:10.1038/ngeo1450 631

GLIMS, National Snow, Ice Data Center, 2005. updated 2012. GLIMS Glacier Database. 632

[Himalaya]. Boulder, Colorado USA: National Snow and Ice Data Center. 633

doi:http://dx.doi.org/10.7265/N5V98602 634

Hambrey, M.J., Quincey, D.J., Glasser, N.F., Reynolds, J.M., Richardson, S.J., Clemmens, 635

S., 2008. Sedimentological, geomorphological and dynamic context of debris-mantled 636

glaciers, Mount Everest (Sagarmatha) region, Nepal. Quaternary Science Reviews 27, 637

2361–2389. doi:10.1016/j.quascirev.2008.08.010 638

Immerzeel, W.W., Kraaijenbrink, P.D.A., Shea, J.M., Shrestha, A.B., Pellicciotti, F., 639

Bierkens, M.F.P., de Jong, S.M., 2013. High-resolution monitoring of Himalayan glacier 640

dynamics using unmanned aerial vehicles. Remote Sensing of Environment 150, 93–103. 641

doi:10.1016/j.rse.2014.04.025 642

Inoue, J., 1977. Mass Budget of Khumbu Glacier: Glaciological Expedition of Nepal, 643

Contribution No. 32. Seppyo 39, 15–19. 644

Inoue, J., Yoshida, M., 1980. Ablation and heat exchange over the Khumbu Glacier. Seppyo 645

41, 26–34. 646

Jouvet, G., Huss, M., Funk, M., Blatter, H., 2011. Modelling the retreat of Grosser 647

Aletschgletscher, Switzerland, in a changing climate. J Glacio 57, 1033–1045. 648

Kääb, A., Berthier, E., Nuth, C., Gardelle, J., Arnaud, Y., 2012. Contrasting patterns of early 649

twenty-first-century glacier mass change in the Himalayas. Nature 488, 495–498. 650

doi:10.1038/nature11324 651

Kessler, M.A., Anderson, R.S., Briner, J.P., 2008. Fjord insertion into continental margins 652

driven by topographic steering of ice. Nature Geosci 1, 365–369. 653

Kirkbride, M.P., Deline, P., 2013. The formation of supraglacial debris covers by primary 654

dispersal from transverse englacial debris bands. Earth Surf. Process. Landforms 38, 655

1779–1792. doi:10.1002/esp.3416 656

Luckman, A., Quincey, D., Bevan, S., 2007. The potential of satellite radar interferometry 657

and feature tracking for monitoring flow rates of Himalayan glaciers. Remote Sensing of 658

Environment 111, 172–181. doi:10.1016/j.rse.2007.05.019 659

Menounos, B., Clague, J.J., Clarke, G.K.C., Marcott, S.A., Osborn, G., Clark, P.U., Tennant, 660

C., Novak, A.M., 2013. Did rock avalanche deposits modulate the late Holocene advance 661

of Tiedemann Glacier, southern Coast Mountains, British Columbia, Canada? Earth and 662

Planetary Science Letters 384, 154–164. doi:10.1016/j.epsl.2013.10.008 663

Mihalcea, C., Mayer, C., Diolaiuti, G., D'Agata, C., Smiraglia, C., Lambrecht, A., 664

Vuillermoz, E., Tartari, G., 2008. Spatial distribution of debris thickness and melting 665

from remote-sensing and meteorological data, at debris-covered Baltoro glacier, 666

Karakoram, Pakistan. Annals of Glaciology 48, 49–57. 667

Moribayashi, S., 1978. Transverse Profiles of Khumbu Glacier Obtained by Gravity 668

Observation: Glaciological Expedition of Nepal, Contribution No. 46. Seppyo 40, 21–25. 669

NASA, 2001. Land Processes Distributed Active Archive Center (LP DAAC). MODIS 670

MOD11C3. USGS/Earth Resources Observation and Science (EROS) Center, Sioux 671

Falls, South Dakota. 672

Nicholson, L., Benn, D., 2006. Calculating ice melt beneath a debris layer using 673

meteorological data. J Glacio 52, 463–470. 674

Nuimura, T., Fujita, K., Yamaguchi, S., Sharma, R.R., 2012. Elevation changes of glaciers 675

revealed by multitemporal digital elevation models calibrated by GPS survey in the 676

Page 22: Modelling the feedbacks between mass balance, ice flow and

21

Khumbu region, Nepal Himalaya, 1992–2008. J Glacio 58, 648–656. 677

doi:10.3189/2012JoG11J061 678

Nye, J.F., 1952. The mechanics of glacier flow. J Glaciol 2, 82–93. 679

Østrem, G., 1959. Ice melting under a thin layer of moraine, and the existence of ice cores in 680

moraine ridges. Geografiska Annaler 41, 228–230. 681

Owen, L.A., Robinson, R., Benn, D.I., Finkel, R.C., Davis, N.K., Yi, C., Putkonen, J., Li, D., 682

Murray, A.S., 2009. Quaternary glaciation of Mount Everest. Quaternary Science 683

Reviews 28, 1412–1433. doi:10.1016/j.quascirev.2009.02.010 684

Pellicciotti, F., Stephan, C., Miles, E., Herreid, S., Immerzeel, W.W., Bolch, T., 2015. Mass-685

balance changes of the debris-covered glaciers in the Langtang Himal, Nepal, from 1974 686

to 1999 61, 1–14. doi:10.3189/2015JoG13J237 687

Quincey, D.J., Luckman, A., Benn, D., 2009. Quantification of Everest region glacier 688

velocities between 1992 and 2002, using satellite radar interferometry and feature 689

tracking. J Glacio 55, 596–606. 690

Ragettli, S., Pellicciotti, F., Immerzeel, W.W., Miles, E.S., Petersen, L., Heynen, M., Shea, 691

J.M., Stumm, D., Joshi, S., Shrestha, A., 2015. Unraveling the hydrology of a Himalayan 692

catchment through integration of high resolution in situ data and remote sensing with an 693

advanced simulation model. Adv Water Resour 78, 94–111. 694

doi:10.1016/j.advwatres.2015.01.013 695

Reid, T.D., Brock, B.W., 2014. Assessing ice-cliff backwasting and its contribution to total 696

ablation of debris-covered Miage glacier, Mont Blanc massif, Italy. J Glacio 60, 3–13. 697

doi:10.3189/2014JoG13J045 698

Reid, T.D., Carenzo, M., Pellicciotti, F., Brock, B.W., 2012. Including debris cover effects in 699

a distributed model of glacier ablation. J. Geophys. Res 117, n/a–n/a. 700

doi:10.1029/2012JD017795 701

Sakai, A., Takeuchi, N., Fujita, K., Nakawo, M., 2000. Role of supraglacial ponds in the 702

ablation process of a debris-covered glacier in the Nepal Himalayas. IAHS Publications 703

119–132. 704

Salerno, F., Guyennon, N., Thakuri, S., Viviano, G., Romano, E., Vuillermoz, E., 705

Cristofanelli, P., Stocchi, P., Agrillo, G., Ma, Y., Tartari, G., 2014. Weak precipitation, 706

warm winters and springs impact glaciers of south slopes of Mt. Everest (central 707

Himalaya) in the last two decades (1994–2013). The Cryosphere Discuss. 8, 5911–5959. 708

doi:10.5194/tcd-8-5911-2014 709

Scherler, D., Bookhagen, B., Strecker, M.R., 2011. Hillslope┽ glacier coupling: The interplay 710

of topography and glacial dynamics in High Asia. J. Geophys. Res 116, F02019. 711

Shea, J.M., Immerzeel, W.W., Wagnon, P., Vincent, C., 2015. Modelling glacier change in 712

the Everest region, Nepal Himalaya. The …. doi:10.5194/tc-9-1105-2015 713

Thakuri, S., Salerno, F., Smiraglia, C., Bolch, T., D'Agata, C., Viviano, G., Tartari, G., 2014. 714

Tracing glacier changes since the 1960s on the south slope of Mt. Everest (central 715

Southern Himalaya) using optical satellite imagery. The Cryosphere 8, 1297–1315. 716

doi:10.5194/tc-8-1297-2014 717

Vacco, D.A., Alley, R.B., Pollard, D., 2010. Glacial advance and stagnation caused by rock 718

avalanches. Earth and Planetary Science Letters 294, 123–130. 719

doi:10.1016/j.epsl.2010.03.019 720

Van Der Veen, C.J., 2013. Fundamentals of glacier dynamics. Second edition. CRS Press. 721

722

723

Table caption 724

Table 1. Simulated glacier volume, ice thickness and velocity during the Little Ice Age (LIA), 725

Page 23: Modelling the feedbacks between mass balance, ice flow and

22

the present day and predicted for AD2100 under a maximum IPCC warming scenario of 726

1.6°C. 727

728

Figure captions 729

Figure 1. Conceptual model of the development of a debris-covered Himalayan glacier; (a) in 730

balance with climate, and (b) during net mass loss under a warming climate. 731

732

Figure 2. Debris-covered Khumbu Glacier. (a) Photograph of the ablation area of Khumbu 733

Glacier looking downglacier from Kala Pathar showing the elevation difference between the 734

Little Ice Age lateral moraine crests and the glacier surface and relict landslide material that 735

has remained in situ from at least 2003 to 2014. (b) Long profile of 100-m mean swath 736

topography of Khumbu Glacier, and (c) long profile of 100-m mean swath topography of 737

Khumbu Glacier below the confluence with the Changri Nup tributary showing the elevation 738

difference between the glacier surface and the Little Ice Age lateral moraine crest due to mass 739

loss by surface lowering. The lowest point of the terminal moraine is 4670 m a.s.l.. 740

741

Figure 3. (a) Present-day ice thickness of Khumbu Glacier estimated from a DEM 742

constructed for the year 2011 as described in Section 3.1, tuned with available ice thickness 743

measurements from radio-echo sounding (Gades et al., 2000) and gravity observations 744

(Moribayashi, 1978), draped over a shaded-relief map of the estimated bedrock topography 745

used to describe the model domain (EBC = Everest Base Camp). (b) Long profile of Khumbu 746

Glacier showing the range of the ELA based on morphometric calculations and 747

measurements of mass balance from previous studies (Benn and Lehmkuhl, 2000; Bolch et 748

al., 2011; Inoue, 1977; Inoue and Yoshida, 1980), the ELA simulated for the present day 749

glacier, and the range of ELA used to represent IPCC scenarios for AD2100. 750

751

Figure 4. Present-day velocities of Khumbu Glacier calculated using feature tracking of the 752

panchromatic bands of multi-temporal Landsat Operational Land Imager (OLI) imagery. The 753

underlying image was acquired by the Landsat OLI on 4th May 2013 and the white areas 754

denote areas with no data. The glacier outline is defined according to the Randolph Glacier 755

Inventory (GLIMS et al., 2005). Note the termination of the measured active ice (i.e. above 756

the uncertainty in the method) is 5.4 km upglacier from the terminus. The location of this 757

figure and the extent of the Central Himalayan region (red dashed line) are shown in the inset 758

map. 759

Page 24: Modelling the feedbacks between mass balance, ice flow and

23

760

Figure 5. Initial steady-state simulation of Khumbu Glacier during the Late Holocene 761

advance (1 ka) used as a starting point for the Little Ice Age simulations, showing (a) ice 762

thickness, (b) debris thickness, (c) mass balance (in metres of water equivalent per year), and 763

(d) velocities. 764

765

Figure 6. Simulations of Khumbu Glacier as a clean-ice rather than debris-covered glacier. 766

(a) present day ice thickness simulated without a supraglacial debris layer, and (b) present-767

day ice thickness simulated without ablation beneath a debris layer with a 50% reduction in 768

ablation. 769

770

Figure 7. Simulations of Khumbu Glacier during the Little Ice Age (LIA; 0.5 ka) and present 771

day. Results from the iSOSIA model for; (a) ice thickness during the LIA, (b) ice thickness at 772

the present day, and simulated supraglacial debris (c) during the LIA, and (d) at the present 773

day. The fit between the simulated glaciers, the LIA lateral moraine crest, and the present day 774

glacier surface are shown for (e) the LIA and (f) the present day. 775

776

Figure 8. Simulated velocities for Khumbu Glacier (a) during the Little Ice Age, and (b) at 777

the present day [Note log scale for velocity], and simulated basal shear stress (IJb) for Khumbu 778

Glacier (c) during the Little Ice Age, and (d) at the present day. 779

780

Figure 9. Simulations of Khumbu Glacier in AD2100 and AD2200. (a) Simulated ice 781

thickness in AD2100 under the maximum IPCC CMIP5 RCP 4.5 warming scenario 782

equivalent to an increase in temperature of 1.6°C from the present day, and (b) assuming the 783

same warming scenario from the present day, the ice thickness in AD2200 after the active 784

glacier detached from the debris-covered tongue. Differences in ice thickness from the 785

present day simulation in (c) AD2100 and (d) AD2200 [Note the different scales for 786

difference in ice thickness]. Debris thickness in (e) AD2100 and (f) AD2200. The Little Ice 787

Age lateral moraine crest, present day glacier surface and simulated glacier surface for (g) the 788

AD2100 and (h) the AD2200 simulations. 789